Skip to main content
Log in

On attitude polarization under Bayesian learning with non-additive beliefs

  • Published:
Journal of Risk and Uncertainty Aims and scope Submit manuscript

Abstract

Ample psychological evidence suggests that people’s learning behavior is often prone to a “myside bias” or “irrational belief persistence” in contrast to learning behavior exclusively based on objective data. In the context of Bayesian learning such a bias may result in diverging posterior beliefs and attitude polarization even if agents receive identical information. Such patterns cannot be explained by the standard model of rational Bayesian learning that implies convergent beliefs. Based on Choquet expected utility theory, we therefore develop formal models of Bayesian learning with psychological bias as alternatives to rational Bayesian learning. We derive conditions under which beliefs may diverge in the learning process despite the fact that all agents observe the same sample drawn from an i.i.d. process. Key to our approach is the description of ambiguous beliefs as neo-additive capacities (Chateauneuf et al., J Econ Theory 137:538–567, 2007), which allows for a flexible and parsimonious parametrization of departures from additive probability measures.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. While agents in our model also apply learning rules by which they will not learn “underlying” probabilities, we motivate the application of these rules by psychological and decision-theoretic arguments.

  2. Related to Doob’s consistency theorem is Blackwell and Dubins’ (1962) convergence theorem for different additive probability measures within the frequentist framework. In particular, Diaconis and Freedman (1986, Theorem 3) establish a formal link between Doob’s consistency theorem and Blackwell and Dubins’ convergence theorem by basically showing that the Bayesian estimate is consistent if and only if any corresponding conditional probability measures merge in the weak topology.

  3. Notice that there are several perceivable definitions of independence for capacities. Very loosely speaking, in the context of conditional capacities Marinacci’s notion of independence corresponds to the optimistic update rule, ensuring that \(\nu \left( A\mid B\right) =\) \(\nu \left( A\right) \) if A and B are independent with respect to the capacity ν.

  4. More precisely, dynamically consistent preferences are at odds with Ellsberg-like behavior whenever the principles of consequentialism and reduction are satisfied, which is the case in Epstein and Schneider’s as well as in our approach (cf. Siniscalchi 2006 and Observation 2 in Zimper 2009).

  5. In the case of learning from ambiguous urns without multiple likelihoods, ambiguity obviously vanishes in the learning process, see Marinacci (2002) for a formal result.

  6. Tonks (1983) introduces a similar model of rational Bayesian learning in which the agent has a normally distributed prior over the mean of some normal distribution and receives normally distributed information.

  7. See Wakker (2009) for a textbook treatment.

References

  • Abel, A. B. (2002). An exploration of the effects of pessimism and doubt on asset returns. Journal of Economic Dynamics and Control, 26, 1075–1092.

    Article  Google Scholar 

  • Acemoglu, D., Chernozhukov, V., & Yildiz, M. (2007). Learning and disagreement in an uncertain world. Working paper, MIT.

  • Alesina, A., Baqir, R., & Easterly, W. (1999). Public goods-and-ethnic divisions. Quarterly Journal of Economics, 115(4), 1167–1199.

    Google Scholar 

  • Baron, J. (2008). Thinking and deciding. New York: Cambridge University Press.

    Google Scholar 

  • Blackwell, D., & Dubins, L. (1962). Merging of opinions with increasing information. The Annals of Mathematical Statistics, 33, 882–886.

    Article  Google Scholar 

  • Brav, A., & Heaton, J. B. (2002). Competing theories of financial anomalies. Review of Financial Studies, 15, 575–606.

    Article  Google Scholar 

  • Breiman, D., LeCam, L., & Schwartz L. (1964). Consistent estimates and zero-one sets. The Annals of Mathematical Statistics, 35, 157–161.

    Article  Google Scholar 

  • Cecchetti, S. G., Lam, P., & Mark, N. C. (2000). Asset pricing with distorted beliefs: Are equity returns too good to be true? American Economic Review, 90, 787–805.

    Article  Google Scholar 

  • Chapman, C. R. (1973). Prior probability bias in information seeking and opinion revision. The American Journal of Psychology, 86, 269–282.

    Article  Google Scholar 

  • Chateauneuf, A., Eichberger, J., & Grant, S. (2007). Choice under uncertainty with the best and worst in mind: Neo-additive capacities. Journal of Economic Theory, 137, 538–567.

    Article  Google Scholar 

  • Diaconis, P., & Freedman, D. (1986). On the consistency of Bayes estimates. The Annals of Statistics, 14, 1–26.

    Article  Google Scholar 

  • Dixit, A. K., & Weibull, J. W. (2007). Political polarization. In Proceedings of the National Academy of Sciences of the United States of America, published online. doi:10.1073/pnas.0702071104.

  • Doob, J. L. (1949). Application of the theory of Martingales. In Le Calcul des Probabilite´s et ses Applications, Colloques Internationaux du Centre National de la Recherche Scientifique (Vol. 13, pp. 23–27).

  • Duclos, J.-Y., Esteban, J., & Ray, D (2004). Polarization: Concepts, measurement, estimation. Econometrica, 72(6), 1737–1772.

    Article  Google Scholar 

  • Eichberger, J., Grant, S., & Kelsey, D. (2006). Updating Choquet beliefs. Journal of Mathematical Economics, 43, 888–899.

    Article  Google Scholar 

  • Eichberger, J., & Kelsey, D. (1999). E-Capacities and the Ellsberg Paradox. Theory and Decision, 46, 107–140.

    Article  Google Scholar 

  • Ellsberg, D. (1961). Risk, ambiguity and the Savage axioms. Quarterly Journal of Economics, 75, 643–669.

    Article  Google Scholar 

  • Epstein, L. G. (1999). A definition of uncertainty aversion. The Review of Economic Studies, 66, 579–608.

    Article  Google Scholar 

  • Epstein, L. G., & Schneider, M. (2003). Recursive multiple-priors. Journal of Economic Theory, 113, 1–31.

    Article  Google Scholar 

  • Epstein, L. G., & Schneider, M. (2007). Learning under ambiguity. Review of Economic Studies, 74, 1275–1303.

    Article  Google Scholar 

  • Epstein, L. G., & Wang, T. (1994). Intertemporal asset pricing under Knightian uncertainty. Econometrica, 62, 283–322.

    Article  Google Scholar 

  • Esteban, J., & Ray, D. (1994). On the measurement of polarization. Econometrica, 62, 819–852.

    Article  Google Scholar 

  • Esteban, J., & Ray, D. (2001). Conflict and distribution. Journal of Economic Theory, 87, 379–415.

    Article  Google Scholar 

  • Fernández, R., & Levy, G. (2008). Diversity and redistribution. Journal of Public Economics, 92(6), 925–943.

    Article  Google Scholar 

  • Freedman, D. A. (1963). On the asymptotic behavior of Bayes’ estimates in the discrete case. The Annals of Mathematical Statistics, 33, 882–886.

    Google Scholar 

  • Fudenberg, D., & Kreps, D. (1993). Learning mixed equilibria. Games and Economic Behavior, 5, 320–367.

    Article  Google Scholar 

  • Fudenberg, D., & Levine, D. K. (1995). Consistency and cautious fictitious play. Journal of Economic Dynamics and Control, 19, 1065–1090.

    Article  Google Scholar 

  • Ghirardato, P., & Marinacci, M. (2002). Ambiguity made precise: A comparative foundation. Journal of Economic Theory, 102, 251–289.

    Article  Google Scholar 

  • Gilboa, I. (1987). Expected utility with purely subjective non-additive probabilities. Journal of Mathematical Economics, 16, 65–88.

    Article  Google Scholar 

  • Gilboa, I., & Schmeidler, D. (1989). Maxmin expected utility with non-unique priors. Journal of Mathematical Economics, 18, 141–153.

    Article  Google Scholar 

  • Gilboa, I., & Schmeidler, D. (1993). Updating ambiguous beliefs. Journal of Economic Theory, 59, 33–49.

    Article  Google Scholar 

  • Kandel, E., & Pearson, N. D. (1995). Differential interpretation of public signals and trade in speculative markets. Journal of Political Economy, 103, 831–872.

    Article  Google Scholar 

  • Krishna, V., & Sjostrom, T. (1998). On the convergence of fictitious play. Mathematics of Operations Research, 23, 479–511.

    Article  Google Scholar 

  • Kurz, M. (1994a). On the structure and diversity of rational beliefs. Economic Theory, 4, 877–900.

    Article  Google Scholar 

  • Kurz, M. (1994b). On rational belief equilibrias. Economic Theory, 4, 859–876.

    Article  Google Scholar 

  • Kurz, M. (1996). Rational beliefs and endogenous uncertainty. Economic Theory, 8, 383–397.

    Google Scholar 

  • Lewellen, J., & Shanken, J. (2002). Learning, asset-pricing tests, and market efficiency. Journal of Finance, 57, 1113–1145.

    Article  Google Scholar 

  • Lijoi, A., Pruenster, I., & Walker, S. O. (2004). Extending Doob’s consistency theorem to nonparametric densities. Bernoulli, 10, 651–663.

    Article  Google Scholar 

  • Lindqvist, E., & Östling, R. (2008). Political polarization and the size of government. SSE/EFI working paper series in economics and finance no. 628.

  • Lord, C. G., Ross, L., & Lepper, M. R. (1979). Biased assimilation and attitude polarization: The effects of prior theories on subsequently considered evidence. Journal of Personality and Social Psychology, 37, 2098–2109.

    Article  Google Scholar 

  • Ludwig, A., & Zimper, A. (2008a). A Parsimonious Choquet model of subjective life-expectancy. Working paper.

  • Ludwig, A., & Zimper, A. (2008b). Biased Bayesian learning and the risk-free rate puzzle. Working paper.

  • Marinacci, M. (1999). Limit laws for non-additive probabilities and their frequentist interpretation. Journal of Economic Literature, 84, 145–195.

    Google Scholar 

  • Marinacci, M. (2002). Learning from ambiguous urns. Statistical Papers, 43, 145–151.

    Article  Google Scholar 

  • Milgrom, P., & Stockey, N. (1982). Information, trade and common knowledge. Journal of Economic Theory, 26, 17–27.

    Article  Google Scholar 

  • Morris, S. (1994). Trade with heterogeneous prior beliefs and asymmetric information. Econometrica, 62, 1327–1347.

    Article  Google Scholar 

  • Neeman, Z. (1996). Approximating agreeing to disagree results with common p-beliefs. Games and Economic Behavior, 16, 77–96.

    Article  Google Scholar 

  • Pires, C. P. (2002). A rule for updating ambiguous beliefs. Theory and Decision, 53, 137–152.

    Article  Google Scholar 

  • Pitz, G. F. (1969). An inertia effect (Resistance to Change) in the revision of opinion. Canadian Journal of Psychology, 23, 24–33.

    Google Scholar 

  • Pitz, G. F., Downing, L., & Reinhold, H. (1967). Sequential effects in the revision of subjective probabilities. Canadian Journal of Psychology, 21, 381–393.

    Google Scholar 

  • Rabin, M., & Schrag, J. L. (1999). First impressions matter: A model of confirmatory bias. Quarterly Journal of Economics, 114(1), 37–82.

    Article  Google Scholar 

  • Rudin, W. (1976). Principles of Mathematical Analysis. New York: McGraw-Hill.

    Google Scholar 

  • Sarin, R., & Wakker, P. P. (1998). Revealed likelihood and Knightian uncertainty. Journal of Risk and Uncertainty, 16, 223–250.

    Article  Google Scholar 

  • Savage, L. J. (1954). The foundations of statistics. New York: Wiley.

    Google Scholar 

  • Schmeidler, D. (1986). Integral representation without additivity. Proceedings of the American Mathematical Society, 97, 255–261.

    Article  Google Scholar 

  • Schmeidler, D. (1989). Subjective probability and expected utility without additivity. Econometrica, 57, 571–587.

    Article  Google Scholar 

  • Siniscalchi, M. (2006). Dynamic choice under ambiguity. Mimeo.

  • Tonks, I. (1983). Bayesian learning and the optimal investment decision of the firm. The Economic Journal, 93, 87–98.

    Article  Google Scholar 

  • Viscusi, W. K. (1985). A Bayesian perspective on biases in risk perception. Economics Letters, 17, 59–62.

    Article  Google Scholar 

  • Viscusi, W. K., & O’Connor, C. J. (1984). Adaptive responses to chemical labeling: Are workers Bayesian decision makers? The American Economic Review, 74 , 942–956.

    Google Scholar 

  • Wakker, P. P. (2001). Testing and characterizing properties of nonadditive measures through violations of the sure-thing principle. Econometrica, 69, 1039–1059.

    Article  Google Scholar 

  • Wakker, P. P. (2004). On the composition of risk preference and belief. Psychological Review, 111, 236–241.

    Article  Google Scholar 

  • Wakker, P. P. (2009). Prospect theory for risk and ambiguity. Cambridge University Press, forthcoming.

  • Weitzman, M. (2007). Subjective expectations and asset-return puzzles. American Economic Review, 97(4), 1102–1130.

    Article  Google Scholar 

  • Zimper, A. (2009). Half-empty, half-full and why we can agree to disagree forever. Journal of Economic Behavior and Organization, 71, 283–299.

    Article  Google Scholar 

Download references

Acknowledgements

We thank Elias Khalil, Robert Östling, Kip Viscusi, Peter Wakker and an anonymous referee for helpful comments and suggestions. Financial support from Economic Research South Africa (ERSA), the German National Research Foundation (DFG) through SFB 504, the State of Baden-Württemberg and the German Insurers Association (GDV) is gratefully acknowledged.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexander Zimper.

Appendix

Appendix

1.1 Formal proofs

Proof of Observation 2

By an argument in Schmeidler (1986, p. 256f), it suffices to restrict attention to a non-negative valued random variable Y so that

$$ E\left[ Y,\nu \right] =\int_{0}^{+\infty }\nu \left\{ \omega \in \Omega \mid Y\left( \omega \right) \geq z\right\} dz\text{,} $$

which is equivalent to

$$ E\left[ Y,\nu \right] =\int_{\min Y}^{\max Y}\nu \left\{ \omega \in \Omega \mid Y\left( \omega \right) \geq z\right\} dz \label{shortCE} $$
(22)

since Y is closed and bounded. We consider a partition P n , n = 1,2,... , of Ω with members

$$ A_{n}^{k}=\left\{ \omega \in \Omega \mid a_{k,n}<X\left( \omega \right) \leq b_{k,n}\right\} \text{ for }k=1,...,2^{n} $$

such that

$$ \begin{array}{rcl} a_{k,n} &=&\left[ \max Y-\min Y\right] \cdot \frac{\left( k-1\right) }{2^{n}} +\min Y \\ b_{k,n} &=&\left[ \max Y-\min Y\right] \cdot \frac{k}{2^{n}}+\min Y\text{.} \end{array} $$

Define the step functions a n :Ω→ℝ and b n :Ω→ℝ such that, for \(\omega \in A_{n}^{k}\), k = 1,...,2n,

$$ \begin{array}{rcl} a_{n}\left( \omega \right) &=&a_{k,n} \\ b_{n}\left( \omega \right) &=&b_{k,n}\text{.} \end{array} $$

Obviously,

$$ E\left[ a_{n},\nu \right] \leq E\left[ Y,\nu \right] \leq E\left[ b_{n},\nu \right] $$

for all n and

$$ \lim_{n\rightarrow \infty }E\left[ b_{n},\nu \right] -E\left[ a_{n},\nu \right] =0\text{.} $$

That is, \(E\left[ a_{n},\nu \right] \) and \(E\left[ b_{n},\nu \right] \) converge to \(E\left[ Y,\nu \right] \) for n→ ∞. Furthermore, observe that

$$ \begin{array}{rcl} \min a_{n} &=&\min Y\text{ for all }n\text{, and} \\ \max b_{n} &=&\max Y\text{ for all }n\text{.} \end{array} $$

Since limn→ ∞  min b n  = limn→ ∞  min a n and \(E\left[ b_{n},\mu \right] \) is continuous in n, we have

$$ \begin{array}{rcl} \lim_{n\rightarrow \infty }E\left[ b_{n},\nu \right] \!&=&\!\delta \left( \lambda \lim_{n\rightarrow \infty }\max b_{n}\!+\!\left( 1\!-\!\lambda \right) \lim_{n\rightarrow \infty }\min b_{n}\right) \!+\!\left( 1\!-\!\delta \right) \lim_{n\rightarrow \infty }E\left[ b_{n},\mu \right] \\ &=&\delta \left( \lambda \max Y+\left( 1-\lambda \right) \min Y\right) +\left( 1-\delta \right) E\left[ Y,\mu \right] \text{.} \end{array} $$

In order to prove Proposition 3, it therefore remains to be shown that, for all n,

$$ E\left[ b_{n},\nu \right] =\delta \left( \lambda \max b_{n}+\left( 1-\lambda \right) \min b_{n}\right) +\left( 1-\delta \right) E\left[ b_{n},\mu \right] \text{.} $$

Since b n is a step function, Eq. 22 becomes

$$ \begin{array}{rcl} E\left[ b_{n},\nu \right] &=&\sum\limits_{A_{n}^{k}\in P_{n}}\nu \left( A_{n}^{2^{n}}\cup ...\cup A_{n}^{k}\right) \cdot \left( b_{k,n}-b_{k-1,n}\right) \\ &=&\sum\limits_{A_{n}^{k}\in P_{n}}b_{k,n}\cdot \left[ \nu \left( A_{n}^{2^{n}}\cup ...\cup A_{n}^{k}\right) -\nu \left( A_{n}^{2^{n}}\cup ...\cup A_{n}^{k-1}\right) \right] \text{,} \end{array} $$

implying for a neo-additive capacity

$$ \begin{array}{rcl} E\left[ b_{n},\nu \right] &=&\max b_{n}\left[ \delta \lambda +\left( 1-\delta \right) \mu \left( A_{n}^{2^{n}}\right) \right] + \sum\limits_{k=2}^{2^{n}-1}b_{k,n}\left( 1-\delta \right) \mu \left( A_{n}^{k}\right) \\ &&+\min b_{n}\left[ 1-\delta \lambda -\left( 1-\delta \right) \sum\limits_{k=2}^{2^{n}}\mu \left( A_{n}^{k}\right) \right] \\ &=&\delta \lambda \max b_{n}+\left( 1-\delta \right) \sum\limits_{k=1}^{2^{n}}b_{k,n}\mu \left( A_{n}^{k}\right) +\min b_{n}\left[ \delta -\delta \lambda \right] \\ &=&\delta \left( \lambda \max b_{n}+\left( 1-\lambda \right) \min b_{n}\right) +\left( 1-\delta \right) E\left[ b_{n},\mu \right] \text{.} \end{array} $$

Proof of Observation 3

An application of the full Bayesian update rule to a neo-additive capacity gives

$$ \begin{array}{rcl} \nu ^{FB}\left( A\mid B\right) &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) +1-\left( \delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cup \lnot B\right) \right) } \\ &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{1+\left( 1-\delta \right) \cdot \left( \mu \left( A\cap B\right) -\mu \left( A\cup \lnot B\right) \right) } \\ &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{1+\left( 1-\delta \right) \cdot \left( \mu \left( A\cap B\right) -\mu \left( A\right) -\mu \left( \lnot B\right) +\mu \left( A\cap \lnot B\right) \right) } \\ &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{1+\left( 1-\delta \right) \cdot \left( -\mu \left( \lnot B\right) \right) } \\ &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{\delta +\left( 1-\delta \right) \cdot \mu \left( B\right) } \\ &=&\frac{\delta \cdot \lambda }{\delta +\left( 1-\delta \right) \cdot \mu \left( B\right) }+\frac{\left( 1-\delta \right) \cdot \mu \left( B\right) }{ \delta +\left( 1-\delta \right) \cdot \mu \left( B\right) }\mu \left( A\mid B\right) \\ &=&\delta _{B}^{FB}\cdot \lambda +\left( 1-\delta _{B}^{FB}\right) \cdot \mu \left( A\mid B\right) \end{array} $$

with \(\delta _{B}^{FB}\) given by Eq. 10. □

Proof of Observation 4

An application of the optimistic Bayesian update rule to a neo-additive capacity gives

$$ \begin{array}{rcl} \nu ^{opt}\left( A\mid B\right) &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cap B\right) }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( B\right) } \\ &=&\frac{\delta \cdot \lambda }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( B\right) }+\frac{\left( 1-\delta \right) \cdot \mu \left( B\right) }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( B\right) }\cdot \mu \left( A\mid B\right) \\[4pt] &=&\delta _{B}^{opt}+\left( 1-\delta _{B}^{opt}\right) \cdot \mu \left( A\mid B\right) \end{array} $$

whereby

$$ \delta _{B}^{opt}=\frac{\delta \cdot \lambda }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( B\right) }\text{.} $$

Proof of Observation 5

An application of the pessimistic Bayesian update rule to a neo-additive capacity gives

$$ \begin{array}{rcl} \nu ^{pess}\left( A\mid B\right) &=&\frac{\nu \left( A\cup \lnot B\right) -\nu \left( \lnot B\right) }{1-\nu \left( \lnot B\right) } \\[3pt] &=&\frac{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( A\cup \lnot B\right) -\delta \cdot \lambda -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) }{1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) } \\[3pt] &=&\frac{\left( 1-\delta \right) \cdot \mu \left( \lnot \left( \lnot A\cap B\right) \right) -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) }{ 1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) } \\[3pt] &=&\frac{\left( 1-\delta \right) \cdot \left( 1-\mu \left( \lnot A\cap B\right) \right) -\left( 1-\delta \right) \cdot \left( 1-\mu \left( B\right) \right) }{1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) } \\[3pt] &=&\frac{\left( 1-\delta \right) \cdot \left( \mu \left( B\right) -\mu \left( \lnot A\cap B\right) \right) }{1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \left( \mu \left( \lnot B\right) \right) } \\[3pt] &=&\frac{\left( 1-\delta \right) \cdot \left( \mu \left( B\right) -\mu \left( B\right) \mu \left( \lnot A\mid B\right) \right) }{1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \left( \mu \left( \lnot B\right) \right) } \\[3pt] &=&\frac{\left( 1-\delta \right) \cdot \left( \mu \left( B\right) -\mu \left( B\right) \left( 1-\mu \left( A\mid B\right) \right) \right) }{ 1-\delta \cdot \lambda -\left( 1-\delta \right) \cdot \mu \left( \lnot B\right) } \\[3pt] &=&\left( 1-\delta _{B}^{pess}\right) \cdot \mu \left( A\mid B\right) \end{array} $$

whereby

$$ \delta _{B}^{pess}=\frac{\delta \left( 1-\lambda \right) }{\delta \left( 1-\lambda \right) +\left( 1-\delta \right) \cdot \mu \left( B\right) }\text{. } $$

Proof of Observation 6

This observation is an immediate consequence of the following lemma combined with the fact that the additive part of the neo-additive beliefs converges in probability to the true probability π  ∗ , i.e.,

$$ \lim_{n\rightarrow \infty }prob\left( \left\vert E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\mid \mathbf{I}_{n}^{k}\right) \right] -\pi ^{\ast }\right\vert \leq c\right) =1 $$

for some c > 0. The lemma simply uses our results (Observations 3–5) on Bayesian updating of neo-additive capacities in order to derive conditional neo-additive capacities for the special case Eq. 17.

Lemma

Suppose that the agent receives sample information \(\mathbf{I}_{n}^{k}\). Contingent on the applied update rule we obtain the following conditional neo-additive beliefs and posterior estimates about parameter π whereby \(E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\mid \mathbf{I}_{n}^{k}\right) \right] \) is given by Eq. 1.

  1. (i)

    Full Bayesian updating.

    $$ \nu ^{FB}\left( \mathbf{\pi }\mid \mathbf{I}_{n}^{k}\right) =\delta _{ \mathbf{I}_{n}^{k}}^{FB}\lambda +\left( 1-\delta _{\mathbf{I} _{n}^{k}}^{FB}\right) \cdot K_{\alpha +k,\beta +n-k}\pi ^{\alpha +k-1}\left( 1-\pi \right) ^{\beta +n-k-1} $$

    with

    $$ \delta _{\mathbf{I}_{n}^{k}}^{FB}=\frac{\delta }{\delta +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}_{n}^{k}\right) } $$

    so that

    $$ E\left[ \tilde{\pi},\nu ^{FB}\left( \mathbf{\cdot }\mid \mathbf{I} _{n}^{k}\right) \right] =\delta _{\mathbf{I}_{n}^{k}}^{FB}\lambda +\left( 1-\delta _{\mathbf{I}_{n}^{k}}^{FB}\right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\mid \mathbf{I}_{n}^{k}\right) \right] . $$
  2. (ii)

    Optimistic Bayesian updating.

    $$ \nu ^{opt}\left( \mathbf{\pi }\mid \mathbf{I}_{n}^{k}\right) =\delta _{ \mathbf{I}_{n}^{k}}^{opt}+\left( 1-\delta _{\mathbf{I}_{n}^{k}}^{opt}\right) \cdot K_{\alpha +k,\beta +n-k}\pi ^{\alpha +k-1}\left( 1-\pi \right) ^{\beta +n-k-1} $$

    with

    $$ \delta _{\mathbf{I}_{n}^{k}}^{opt}=\frac{\delta \cdot \lambda }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}_{n}^{k}\right) } $$

    so that

    $$ E\left[ \tilde{\pi},\nu ^{opt}\left( \mathbf{\cdot }\mid \mathbf{I} _{n}^{k}\right) \right] =\delta _{\mathbf{I}_{n}^{k}}^{opt}+\left( 1-\delta _{\mathbf{I}_{n}^{k}}^{opt}\right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\mid \mathbf{I}_{n}^{k}\right) \right] . $$
  3. (iii)

    Pessimistic Bayesian updating.

    $$ \nu ^{pess}\left( \mathbf{\pi }\mid \mathbf{I}_{n}^{k}\right) =\left( 1-\delta _{\mathbf{I}_{n}^{k}}^{pess}\right) \cdot K_{\alpha +k,\beta +n-k}\pi ^{\alpha +k-1}\left( 1-\pi \right) ^{\beta +n-k-1} $$

    with

    $$ \delta _{\mathbf{I}_{n}^{k}}^{pess}=\frac{\delta \cdot \left( 1-\lambda \right) }{\delta \cdot \left( 1-\lambda \right) +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}_{n}^{k}\right) } $$

    so that

    $$ E\left[ \tilde{\pi},\nu ^{pess}\left( \mathbf{\cdot }\mid \mathbf{I} _{n}^{k}\right) \right] =\left( 1-\delta _{\mathbf{I}_{n}^{k}}^{pess}\right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\mid \mathbf{I} _{n}^{k}\right) \right] . $$

Proof of Observation 7

At first observe that δ > 0 and \(\lambda \in \left( 0,1\right) \) implies \(\delta _{\mathbf{I} ^{\ast }}^{FB}>\delta _{\mathbf{I}^{\ast }}^{opt}\) as well as \(\delta _{ \mathbf{I}^{\ast }}^{FB}>\delta _{\mathbf{I}^{\ast }}^{pess}\). Consider the inequality

$$ \begin{array}{rcl} E\left[ \tilde{\pi},\nu ^{FB}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] &<&E\left[ \tilde{\pi},\nu ^{opt}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] \Leftrightarrow \\ \delta _{\mathbf{I}^{\ast }}^{FB}\cdot \lambda &<&\delta _{\mathbf{I}^{\ast }}^{opt}+\left( \delta _{\mathbf{I}^{\ast }}^{FB}-\delta _{\mathbf{I}^{\ast }}^{opt}\right) \cdot \pi ^{\ast }\text{,} \end{array} $$

which holds, by \(\delta _{\mathbf{I}^{\ast }}^{FB}>\delta _{\mathbf{I}^{\ast }}^{opt}\), for all π  ∗  iff

$$ \begin{array}{rcl} \delta _{\mathbf{I}^{\ast }}^{FB}\cdot \lambda &<&\delta _{\mathbf{I}^{\ast }}^{opt}\Leftrightarrow \\ \frac{\delta \cdot \lambda }{\delta +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}^{\ast }\right) } &<&\frac{\delta \cdot \lambda }{\delta \cdot \lambda +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}^{\ast }\right) }\Leftrightarrow \\ \lambda &<&1\text{.} \end{array} $$

Turn now to the inequality

$$ \begin{array}{rcl} E\left[ \tilde{\pi},\nu ^{pess}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] &<&E\left[ \tilde{\pi},\nu ^{FB}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] \Leftrightarrow \\ \left( \delta _{\mathbf{I}^{\ast }}^{FB}-\delta _{\mathbf{I}^{\ast }}^{pess}\right) \cdot \pi ^{\ast } &<&\delta _{\mathbf{I}^{\ast }}^{FB}\cdot \lambda \text{,} \end{array} $$

which holds, by \(\delta _{\mathbf{I}^{\ast }}^{FB}>\delta _{\mathbf{I}^{\ast }}^{pess}\), for all π  ∗  iff

$$ \begin{array}{rcl} \left( \delta _{\mathbf{I}^{\ast }}^{FB}-\delta _{\mathbf{I}^{\ast }}^{pess}\right) &<&\delta _{\mathbf{I}^{\ast }}^{FB}\cdot \lambda \Leftrightarrow \\ \delta _{\mathbf{I}^{\ast }}^{FB}\cdot \left( 1-\lambda \right) &<&\delta _{ \mathbf{I}^{\ast }}^{pess}\Leftrightarrow \\ \frac{\delta \cdot \left( 1-\lambda \right) }{\delta +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}^{\ast }\right) } &<&\frac{\delta \cdot \left( 1-\lambda \right) }{\delta \cdot \left( 1-\lambda \right) +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I}^{\ast }\right) }\Leftrightarrow \\ 0 &<&\lambda \text{.} \end{array} $$

This proves the first part of the observation. The second part readily follows from the assumption that \(\mu \left( \mathbf{I}^{\ast }\right) <1\). □

Proof of Proposition 1

Part (i). :

Observe at first that inequality (19) is satisfied if and only if λ 1 ≥ λ 2. Obviously, if λ 1 = λ 2 then Eq. 18 must be violated. Thus we can restrict attention to λ 1 > λ 2. Observe that, by the corollary, Eq. 18 writes as

$$ \begin{array}{rcl} &&\delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{1}+\left( 1-\delta _{\mathbf{I} ^{\ast }}^{FB}\right) \cdot \pi ^{\ast }-\delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{2}+\left( 1-\delta _{\mathbf{I}^{\ast }}^{FB}\right) \cdot \pi ^{\ast } \\ &>&\delta \cdot \lambda _{1}+\left( 1-\delta \right) \cdot E\left[ \tilde{\pi },\mu \left( \mathbf{\cdot }\right) \right] -\left( \delta \cdot \lambda _{2}+\left( 1-\delta \right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{ \cdot }\right) \right] \right) \text{,} \end{array} $$

which is equivalent to

$$ \begin{array}{rcl} \delta _{\mathbf{I}^{\ast }}^{FB} &>&\delta \Leftrightarrow \\ \frac{\delta }{\delta +\left( 1-\delta \right) \cdot \mu \left( \mathbf{I} ^{\ast }\right) } &>&\delta \text{,} \end{array} $$

and therefore holds if and only if \(\delta \in \left( 0,1\right) \) since \( \mu \left( \mathbf{I}^{\ast }\right) <1\).

Part (ii). :

Again, observe at first that inequality (19) is satisfied if and only if λ 1 ≥ λ 2. By the corollary, Eq. 18 becomes

$$ \begin{array}{rcl} &&\delta _{\mathbf{I}^{\ast }}^{opt}+\left( 1-\delta _{\mathbf{I}^{\ast }}^{opt}\right) \cdot \pi ^{\ast }-\left( 1-\delta _{\mathbf{I}^{\ast }}^{pess}\right) \cdot \pi ^{\ast } \\ &>&\delta \cdot \lambda _{1}+\left( 1-\delta \right) \cdot E\left[ \tilde{\pi },\mu \left( \mathbf{\cdot }\right) \right] -\left( \delta \cdot \lambda _{2}+\left( 1-\delta \right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{ \cdot }\right) \right] \right) \end{array} $$

which is equivalent to

$$ \delta _{\mathbf{I}^{\ast }}^{opt}+\left( \delta _{\mathbf{I}^{\ast }}^{pess}-\delta _{\mathbf{I}^{\ast }}^{opt}\right) \cdot \pi ^{\ast }>\delta \left( \lambda _{1}-\lambda _{2}\right) . \label{inequ} $$
(23)

If δ = 0, the l.h.s. as well as the r.h.s. of Eq. 23 equal zero. Thus, δ > 0 is a necessary condition for Eq. 23 to hold. In what follows we prove that δ > 0 is also a sufficient condition. Let δ > 0 and consider at first the case that

$$ \delta _{\mathbf{I}^{\ast }}^{pess}-\delta _{\mathbf{I}^{\ast }}^{opt}\leq 0 \text{.} \label{case1} $$
(24)

Since the l.h.s. of Eq. 23 is then continuously strictly decreasing in π  ∗  and, by assumption, \(\pi ^{\ast }\in \left( 0,1\right) \), Eq. 23 is satisfied for all π  ∗  if and only if

$$ \begin{array}{rl} \delta _{\mathbf{I}^{\ast }}^{pess}& \geq \delta \left( \lambda _{1}-\lambda _{2}\right) \Leftrightarrow \\ \frac{1-\lambda _{2}}{\delta (1-\lambda _{2})+(1-\delta )\mu (\mathbf{I} ^{\ast })}& \geq \lambda _{1}-\lambda _{2}\text{,} \end{array} $$

which is obviously true for all λ 1,λ 2 since

$$ \frac{1-\lambda _{2}}{\delta (1-\lambda _{2})+(1-\delta )\mu (\mathbf{I} ^{\ast })}\geq 1-\lambda _{2}\text{.} $$

This proves the claim for case (24).

Let δ > 0 and consider now the converse case

$$ \delta _{\mathbf{I}^{\ast }}^{pess}-\delta _{\mathbf{I}^{\ast }}^{opt}>0 \text{.} \label{case2} $$
(25)

Since the l.h.s. of Eq. 23 is then continuously strictly increasing in \(\pi ^{\ast }\in \left( 0,1\right) \), Eq. 23 is satisfied for all π  ∗  if and only if

$$ \begin{array}{rcl} \delta _{\mathbf{I}^{\ast }}^{opt}+\left( \delta _{\mathbf{I}^{\ast }}^{pess}-\delta _{\mathbf{I}^{\ast }}^{opt}\right) &\geq &\delta \left( \lambda _{1}-\lambda _{2}\right) \Leftrightarrow \\[3pt] \delta _{\mathbf{I}^{\ast }}^{pess} &\geq &\delta \left( \lambda _{1}-\lambda _{2}\right) \Leftrightarrow \\[3pt] \frac{1-\lambda _{2}}{\delta (1-\lambda _{2})+(1-\delta )\mu (\mathbf{I} ^{\ast })} &\geq &\lambda _{1}-\lambda _{2}\text{,} \end{array} $$

which is obviously true for all λ 1,λ 2. This proves that δ > 0 is sufficient for Eq. 23 to hold. □

Proof of Proposition 2

Part (i). :

By the corollary and the assumption that \(E\left[ \tilde{ \pi},\mu \left( \mathbf{\cdot }\right) \right] =\pi ^{\ast }\) Eq. 20 implies

$$ \begin{array}{rl} E\left[ \tilde{\pi},\nu _{1}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] &> E\left[ \tilde{\pi},\nu _{2}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] \Leftrightarrow \\ \delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{1}+\left( 1-\delta _{\mathbf{I} ^{\ast }}^{FB}\right) \cdot \pi ^{\ast } & >\delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{2}+\left( 1-\delta _{\mathbf{I}^{\ast }}^{FB}\right) \cdot \pi ^{\ast } \end{array} $$

which holds if and only if λ 1 > λ 2 so that the middle inequality in Eq. 20 is also strict. Focus now on the inequalities

$$ \begin{array}{rl} E\left[ \tilde{\pi},\nu _{1}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] &>E\left[ \tilde{\pi},\nu _{1}\left( \mathbf{\cdot }\right) \right] \Leftrightarrow \\ \delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{1}+\left( 1-\delta _{\mathbf{I} ^{\ast }}^{FB}\right) \cdot \pi ^{\ast } &>\delta \cdot \lambda _{1}+\left( 1-\delta \right) \cdot \pi ^{\ast } \end{array} $$

and

$$ \begin{array}{rl} E\left[ \tilde{\pi},\nu _{2}\left( \mathbf{\cdot }\right) \right] & >E\left[ \tilde{\pi},\nu _{2}\left( \mathbf{\cdot }\mid \mathbf{I}^{\ast }\right) \right] \Leftrightarrow \\ \delta \cdot \lambda _{2}+\left( 1-\delta \right) \cdot \pi ^{\ast } &>\delta _{\mathbf{I}^{\ast }}^{FB}\lambda _{2}+\left( 1-\delta _{\mathbf{I} ^{\ast }}^{FB}\right) \cdot \pi ^{\ast } \end{array} $$

which are implied by Eq. 20 under the assumption that \(E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\right) \right] =\pi ^{\ast }\). Observe that these inequalities require \(\delta \in \left( 0,1\right) \) since \(\delta \in \left\{ 0,1\right\} \) would imply \(\delta _{\mathbf{I} ^{\ast }}^{FB}=\delta \). As a consequence of \(\delta \in \left( 0,1\right)\), we have from the corollary that \(\delta _{\mathbf{I}^{\ast }}^{FB}>\delta \) because \(\mu \left( \mathbf{I}^{\ast }\right) <1\) so that the above inequalities hold if and only if

$$ \lambda _{1}>\pi ^{\ast }>\lambda _{2}\text{,} $$

which proves the result.

Part (ii). :

By the corollary, the inequality \(E\left[ \tilde{\pi} ,\nu _{1}\left( \mathbf{\cdot }\right) \right] \geq E\left[ \tilde{\pi},\nu _{2}\left( \mathbf{\cdot }\right) \right] \) in Eq. 20 holds if and only if λ 1 ≥ λ 2. Consider at first agent 1 and rewrite the relevant part in Eq. 20 as

$$ \begin{array}{rcl} E\left[ \tilde{\pi},\nu _{1}^{opt}\left( \mathbf{\cdot }\mid \mathbf{I} ^{\ast }\right) \right] &>&E\left[ \tilde{\pi},\nu _{1}\left( \mathbf{\cdot } \right) \right] \Leftrightarrow \\ \delta _{\mathbf{I}^{\ast }}^{opt}+\left( 1-\delta _{\mathbf{I}^{\ast }}^{opt}\right) \cdot \pi ^{\ast } &>&\delta \cdot \lambda _{1}+\left( 1-\delta \right) \cdot E\left[ \tilde{\pi},\mu \left( \mathbf{\cdot }\right) \right] \end{array} $$

which, under the assumption that \(E\left[ \tilde{\pi},\mu \left( \mathbf{ \cdot }\right) \right] =\pi ^{\ast }\) is equivalent to

$$ \delta _{\mathbf{I}^{\ast }}^{opt}+\left( \delta -\delta _{\mathbf{I}^{\ast }}^{opt}\right) \cdot \pi ^{\ast }>\delta \cdot \lambda _{1}. \label{inequal2} $$
(26)

Observe that δ > 0 is a necessary condition for Eq. 26 to hold. In what follows we prove that δ > 0 is also sufficient. Let δ > 0 and consider at first the case that

$$ \delta -\delta _{\mathbf{I}^{\ast }}^{opt}\leq 0\text{.} \label{case3} $$
(27)

Since the l.h.s. of Eq. 26 is then continuously strictly decreasing in \(\pi ^{\ast }\in \left( 0,1\right) \), Eq. 26 is satisfied for all π  ∗  if and only if

$$ \begin{array}{rcl} \delta _{\mathbf{I}^{\ast }}^{opt}& \geq &\delta \cdot \lambda _{1}\Leftrightarrow \\ \frac{1}{\delta \lambda _{1}+\left( 1-\delta \right) \mu (\mathbf{I}^{\ast }) }& \geq &1\text{,} \end{array} $$

which is obviously satisfied. This proves the claim for case (27).

Let δ > 0 and consider now the converse case

$$ \delta -\delta _{\mathbf{I}^{\ast }}^{opt}>0\text{.} \label{case4} $$
(28)

Since the l.h.s. of Eq. 26 is then continuously strictly increasing in \(\pi ^{\ast }\in \left( 0,1\right) \), Eq. 26 is satisfied for all π  ∗  if and only if

$$ \delta \geq \delta \cdot \lambda _{1}\text{,} $$

which is obviously satisfied. This proves our claim that δ > 0 is a necessary and sufficient condition for Eq. 26 to hold. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Zimper, A., Ludwig, A. On attitude polarization under Bayesian learning with non-additive beliefs. J Risk Uncertain 39, 181–212 (2009). https://doi.org/10.1007/s11166-009-9074-0

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11166-009-9074-0

Keywords

JEL Classification

Navigation